Development of variously functionalized nitrile oxides Haruyasu Asahara*1,2, Keita Arikiyo1 and Nagatoshi Nishiwaki*1,2

Full Research Paper Address: 1School of Environmental Science and Engineering, Kochi University of Technology, Tosayamada, Kami, Kochi 782-8502, Japan and 2Research Center for Material Science and Engineering, Kochi University of Technology, Tosayamada, Kami, Kochi 782-8502, Japan Email: Haruyasu Asahara* - [email protected]; Nagatoshi Nishiwaki* - [email protected] * Corresponding author

Open Access Beilstein J. Org. Chem. 2015, 11, 1241–1245. doi:10.3762/bjoc.11.138 Received: 13 May 2015 Accepted: 16 July 2015 Published: 23 July 2015 Associate Editor: R. Sarpong © 2015 Asahara et al; licensee Beilstein-Institut. License and terms: see end of document.

Keywords: acylnitrile oxide; amide; formylnitrile oxide; functionalized nitrile oxide; Weinreb amide

Abstract N-Methylated amides (N,4-dimethylbenzamide and N-methylcyclohexanecarboxamide) were systematically subjected to chemical transformations, namely, N-tosylation followed by nucleophilic substitution. The amide function was converted to the corresponding carboxylic acid, esters, amides, aldehyde, and ketone upon treatment with hydroxide, alkoxide, amine, diisobutylaluminium hydride and Grignard reagent, respectively. In these transformations, N-methyl-N-tosylcarboxamides behave like a Weinreb amide. Similarly, N-methyl-5-phenylisoxazole-3-carboxamide was converted into 3-functionalized isoxazole derivatives. Since the amide was prepared by the cycloaddition reaction of ethynylbenzene and N-methylcarbamoylnitrile oxide, the nitrile oxide served as the equivalent of the nitrile oxides bearing a variety of functional groups such as carboxy, alkoxycarbonyl, carbamoyl, acyl and formyl moieties.

Introduction Nitrile oxides are valuable synthetic synthons for the construction of heterocyclic compounds by cycloaddition reactions, which lead to the formation of two bonds in a single experimental step [1,2]. In addition, cycloadducts serve as precursors of polyfunctionalized compounds by ring opening reaction followed by N–O bond fission [2,3]. Hence, functionalized nitrile oxides are clearly useful for the preparation of more complex systems. However, because the precursors of the functionalized nitrile oxides are not always easily available, there are only a few reports that deal with these compounds compared to those on alkylated and arylated nitrile oxides [1]. Thus, it is highly

desirable to develop methodologies for the facile generation of functionalized nitrile oxides. In our previous work, we reported the preparation of nitrile oxide 2 bearing an N-methylcarbamoyl group by treatment of 2-methyl-4-nitroisoxazolin-5(2H)-one (1) only with water at room temperature. Nitrile oxide 2 undergoes cycloaddition reactions with alkenes [4,5], alkynes (Scheme 1) [4,5], nitriles [6], and 1,3-dicarbonyl compounds [7] to afford the corresponding polyfunctionalized heterocyclic compounds, which are not easily obtained by other approaches. Although this protocol is

1241

Beilstein J. Org. Chem. 2015, 11, 1241–1245.

Scheme 1: A strategy for developing functionalized nitrile oxides.

expected to be useful for synthesizing polyfunctionalized compounds, it is limited by the need of a N-methylcarbamoyl functional group. The synthetic utility of nitrile oxide 2 will be indeed improved by converting the N-methylcarbamoyl group into other functionalities, i.e., nitrile oxide 2 may serve as an equivalent of nitrile oxides 4 having versatile functional groups (Scheme 1).

already reported by Itoh and coworkers [10]. However, a systematic study using simple amides has not yet been performed. In this context, we examined the sulfonylation and chemical conversion of N-methylated aromatic and aliphatic amides, and a similar conversion of N-methyl-5-phenylisoxazole-3-carboxamide (3), which is equal to the generation of variously functionalized nitrile oxides.

Weinreb amide (N-methoxy-N-methylamide) is widely used for the conversion of less reactive amide moieties into other carbonyl functionalities [8,9]. The two oxygen atoms of the methoxy and carbonyl groups coordinate to the organometallic reagents, thereby suppressing further reaction with the obtained aldehyde or ketone. However, this protocol cannot be applied to the conversion of N-methylamides because methoxylation of N-methylamides is hitherto unknown. Thus, we speculated that the introduction of a sulfonyl group would be effective because of its high electron-withdrawing and coordinating abilities, serving as an equivalent of the Weinreb amide. Indeed, the use of N-methyl-3-phenyl-N-tosylbutanamide for this purpose was

Results and Discussion At the outset, methanesulfonylation (mesylation) of amides was studied. To a suspension of N,4-dimethylbenzamide (6A) and sodium hydride (2 equiv) in THF, mesyl chloride (7a, 1.3 equiv) was added, and the resultant mixture was stirred at room temperature for 9 h. After acidification, the reaction mixture was extracted with diethyl ether to afford the crude material, from which 32% of N-mesylated product 8Aa [11] was isolated and 61% of 6A was recovered. Thus, the conversion yield of 8Aa was 82% (Table 1, entry 1). On the other hand, the yield of 8Aa was considerably increased (up to 70%) when ethylmagnesium bromide was employed as a base instead of

Table 1: Sulfonylation of N-methylamides.

entry

R1

R2

Base (equiv)

Temp./°C

Time/h

Product

Yield/%

1 2 3 4 5b

4-MeC6H4 4-MeC6H4 4-MeC6H4 c-Hexyl c-Hexyl

Me Me 4-MeC6H4 4-MeC6H4 4-MeC6H4

NaH (2) EtMgBr (1.1) NaH (2) NaH (2) NaH (2)

rt 0 0 0 0

9 1 24 12 12

8Aa 8Aa 8Ab 8Bb 8Bb

32 (82)a 70 60 (92)a 38 60 (85)a

aConversion

yields were shown in parentheses; b5 equivalents of 7b were used.

1242

Beilstein J. Org. Chem. 2015, 11, 1241–1245.

sodium hydride (Table 1, entry 2). However, the instability and gradual decomposition of mesylated product 8Aa by ambient moisture impedes its storage and use as a synthetic reagent. Hence, in the place of the mesyl group, the 4-methylbenzenesulfonyl (tosyl) moiety was employed as the activating group of the amide function. Indeed, tosylated product 8Ab [11] proved to be sufficiently stable under ambient conditions. Despite the considerable efforts for optimizing the reaction conditions, the highest yield obtained for 8Ab was 60%; however, 35% of starting material 6A was recovered, and the conversion yield was 92%. This tosylation method was applied to the aliphatic amide N-methylcyclohexanecarboxamide (6B) to afford 8Bb upon treatment with tosyl chloride (7b) using sodium hydride as a base (Table 1, entry 4). In this case, the use of an excess of 7b increased the yield of 8Bb up to 60% (85% of conversion yield, Table 1, entry 5). Next, nucleophilic substitutions of tosylated amides 8Ab and 8Bb were investigated (Table 2). Whereas water and methanol did not alter the amide moieties, hydroxide and methoxide converted amides 8Ab and 8Bb into the corresponding carboxylic acids (9a and 10a) and methyl esters (9b and 10b),

respectively (Table 2, entries 1–4). In addition, the nucleophilic substitution also proceeded by employing neutral amines. The reaction of 8 with propylamine proceed smoothly, leading to N-propylamides 9c and 10c (Table 2, entries 5 and 6). The substitution reaction was influenced by the bulkiness of the amines. Heating at 120 °C was required for the substitution reaction by sec-butylamine. However, in the case of tert-butylamine, even when the reaction was conducted in a sealed tube at 120 °C, only a small amount of amide was observed (Table 2, entries 7–10). On the other hand, a cyclic secondary amine, pyrrolidine, afforded amides 9f and 10f quantitatively at room temperature (Table 2, entries 11 and 12). When more reactive butyllithium or isopropylmagnesium bromide was used as the nucleophile, the corresponding tertiary alcohols 11g and 11h were obtained in 51% and 58% yields, respectively, whereas ketones 9g and 9h were not detected (Table 2, entries 13 and 14). In the case of an aromatic Grignard reagent, the undesired reaction was suppressed. Although the reaction did not occur at −78 °C, the substitution proceeded successfully at −40 °C, affording ketones 9i and 10i (Table 2, entries 15 and 16). Sodium borohydride also proved to be

Table 2: Chemical conversions of N-methylamide to other functions.

entry

Amide

Nucleophile

Solvent

Temp./°C

Product

Yield/%

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

8Ab 8Bb 8Ab 8Bb 8Ab 8Bb 8Ab 8Bb 8Ab 8Bb 8Ab 8Bb 8Ab 8Ab 8Ab 8Bb 8Ab 8Ab 8Bb

NaOH NaOH NaOMe NaOMe PrNH2 PrNH2 sec-BuNH2 sec-BuNH2 tert-BuNH2 tert-BuNH2 pyrrolidine pyrrolidine BuLi iPrMgBr 4-MeOC6H4MgBr 4-MeOC6H4MgBr NaBH4 iBu2AlH iBu2AlH

H2O H2O MeOH MeOH THF THF THF THF THF THF THF THF THF THF THF THF THF THF THF

85 85 60 60 rt rt 120 120 120 120 rt rt −78 −40 −40 −40 −78 −78 −78

9a 10a 9b 10b 9c [12] 10c 9d [13] 10d 9e [14] 10e [15] 9f [16] 10f [17] 9g 9h 9i [18] 10i [19] 9j 9j 10j

quant. 94 62 73 quant. 82 72 79 19 3 quant. quant. 0 (51)a 0 (58)a 59 33 0 (64)a 77 (12)a 39

aIntermediately

produced 9 or 10 underwent the excess reactions to afford alcohols 11.

1243

Beilstein J. Org. Chem. 2015, 11, 1241–1245.

highly reactive, causing a further addition reaction even at −78 °C to generate the corresponding alcohol 11j (Table 2, entry 17). This disadvantage was overcome by using bulkier diisobutylaluminium hydride (DIBAL) to furnish aldehydes 9j and 10j, although small amounts of alcohol 11j were also detected (Table 2, entries 18 and 19). In contrast to tosylated amides 8Ab and 8Bb, tosylated N-methyl-5-phenylisoxazole-3-carboxamide 12 exhibited higher reactivity. Upon tosylation under similar conditions 3 did not afford product 12, and isoxazole-3-carboxylic acid 5a, a hydrolyzed product of 12, was observed. This problem was solved by quenching the reaction mixture with ice water instead of just water; yet, 12 was afforded in only 10% yield, and 5a was the main product in 38% yield. These findings imply that during the work-up procedure, further decomposition of 12 by water occurred even at low temperatures because of the higher electron deficiency of the isoxazole ring compared with carbocyclic rings. When, instead of water, less nucleophilic acetic acid was employed for the quench, 12 was isolated in 74% yield and the competitive hydrolysis of 12 was suppressed (6% of 5a). Although tosylated amide 12 could be isolated, it gradually decomposed under ambient conditions. Hence, the chemical conversion was performed by adding a nucleophile to the mixture of the tosylation reaction, without the isolation of 12 (Table 3). As mentioned above, the use of a hydroxide anion for the hydrolysis was not required, and carboxylic acid 5a was obtained in 81% yield by the addition of only water (Table 3, entry 1). This protocol was applicable to a relatively bulky alcohol to afford isopropyl ester 5k in moderate yield (Table 3, entry 2). Similarly, primary and secondary amines underwent the substitution reaction, leading to 5c–f in good to moderate yields (Table 3, entries 3–6). It was noteworthy that, because of high reactivity of 12 caused by the isoxazole ring [20], the methylamino group could be replaced by a bulky tert-butylamino moiety. Conversion of the N-methylcarbamoyl group to an acyl or formyl group was achieved upon treatment with a Grignard reagent or with DIBAL (Table 3, entries 7 and 8). The chemical conversion of N-methylamides to versatile carbonyl functions was systematically studied. In this protocol, the tosyl group was found to be effective for the activation of the carbamoyl group and for preventing over-addition by organometallic reagents. Isoxazole-3-carboxamide 3 was successfully converted in a similar fashion. Thus, as shown in Figure 1, nitrile oxide 2 serves as an equivalent of functionalized nitrile oxides 4, thereby improving the synthetic utility of carbamoylnitrile oxide 2.

Table 3: Chemical conversion of N-methylisoxazole-3-carboxamide 3.

entry

Nu-H

1 2 3 4 5 6 7a,b

OH2 iPrOH PrNH2 sec-BuNH2 tert-BuNH2 pyrrolidine 4-MeOC6H4 MgBr iBu2AlH

8a,c a2

Equiv of nucleophiles were used.

Yield/% 5a [21] 5k [22] 5c [23] 5d [22] 5e [24] 5f [23] 5i [25]

81 59 74 59 30 42 45

5j [26] bAt

−40 °C.

40 cAt

−78°C.

Figure 1: Versatile functionalized nitrile oxides.

Experimental Conversion of isoxazolecarboxamide 3 to 5c via tosyl derivative 12 To a solution of amide 3 (101 mg, 0.5 mmol) in THF (1 mL), a suspension of 60 wt % sodium hydride (100 mg, 2.5 mmol) in THF (3 mL) was added under argon. After the mixture was stirred vigorously for 10 min, the mixture was cooled to 0 °C and then a solution of tosyl chloride (191 mg, 1 mmol) in THF (1 mL) was slowly added. After the mixture was stirred for 3 h at 0 °C, propylamine (164 μL, 2 mmol) was added, and then the mixture was stirred at room temperature for 12 h. After evapor-

1244

Beilstein J. Org. Chem. 2015, 11, 1241–1245.

ation of the solvent, the residue was dissolved into diethyl ether (5 mL), washed with water (5 mL), and the aqueous layer was extracted with diethyl ether (2 × 5 mL). The combined organic layer was dried over magnesium sulfate, and concentrated, and the residue was subjected to column chromatography on silica gel to afford N-propylcarboxamide 5c (85 mg, 0.37 mol, 74% yield), eluted with dichloromethane. 1 H NMR (400 MHz, CDCl 3 , TMS) δ 1.01 (t, J = 7.2 Hz, 3H), 1.67 (tq, J = 7.2, 7.2 Hz, 2H), 3.43 (dt, J = 7.2, 7.2 Hz, 2H), 6.79–6.90 (br, 1H), 6.97 (s, 1H), 7.47–7.50 (m, 3H), 7.78–7.81 (m, 2H); 13C NMR (100 MHz, CDCl 3 , TMS) δ 11.4 (CH 3 ), 22.7 (CH 2 ), 41.2 (CH2), 99.2 (CH), 126.0 (CH), 126.9 (C), 129.1 (CH), 130.7 (CH), 158.9 (C), 159.3 (C), 171.5 (C).

References 1. Nishiwaki, N.; Asahara, H. Carbamoylnitrile Oxide and Inverse Electron-Demand 1,3-Dipolar Cycloaddition. In Methods and Applications of Cycloaddition Reactions in Organic Syntheses; Nishiwaki, N., Ed.; John Wiley & Sons: Hoboken, USA, 2014; pp 223–240. doi:10.1002/9781118778173.ch09 2. Torssell, K. B. G. Nitrile Oxides, Nitrones, and Nitronates in Organic Synthesis; VCH Publishers, Inc.: New York, USA, 1988; pp 55–74.

16. Kovalenko, O. O.; Volkov, A.; Adolfsson, H. Org. Lett. 2015, 17, 446–449. doi:10.1021/ol503430t 17. Kawano, M.; Kiuchi, T.; Matsuo, J.-i.; Ishibashi, H. Tetrahedron Lett. 2012, 53, 432–434. doi:10.1016/j.tetlet.2011.11.067 18. Skillinghaug, B.; Sköld, C.; Rydfjord, J.; Svensson, F.; Behrends, M.; Sävmarker, J.; Sjöberg, P. J. R.; Larhed, M. J. Org. Chem. 2014, 79, 12018–12032. doi:10.1021/jo501875n 19. Shivakumar, S. B.; Basavaraju, Y. B.; Umesha, B.; Krishna, M. H.; Mallesha, N. Eur. J. Chem. 2014, 5, 424–429. doi:10.5155/eurjchem.5.3.424-429.1020 20. Tamura, M.; Nishimura, T.; Nishiwaki, N.; Ariga, M. Heterocycles 2004, 63, 1659–1665. doi:10.3987/COM-04-10085 21. Desai, V. G.; Naik, S. R.; Dhumaskar, K. L. Synth. Commun. 2014, 44, 1453–1460. doi:10.1080/00397911.2013.854916 22. Commercially available. 23. Tait, B.; Cullen, M. Methods of modulating cftr activity. PCT Int. Appl. WO 2014210159, Dec 31, 2014. 24. Maywald, V.; Freund, W.; Hamprecht, G.; Kuekenhoehner, T.; Plath, P.; Wuerzer, B.; Westphalen, K.-O. Carbonsäureamide. Eur. Pat. Appl. EP 418667, March 27, 1991. 25. Gao, P.; Li, H.-X.; Hao, X.-H.; Jin, D.-P.; Chen, D.-Q.; Yan, X.-B.; Wu, X.-X.; Song, X.-R.; Liu, X.-Y.; Liang, Y.-M. Org. Lett. 2014, 16, 6298–6301. doi:10.1021/ol503228x 26. Dikusar, E. A.; Potkin, V. I.; Pyatkevich, S. K. Russ. J. Gen. Chem. 2014, 84, 1701–1705. doi:10.1134/S1070363214090102

3. Nishiwaki, N.; Maki, A.; Ariga, M. J. Oleo Sci. 2009, 58, 481–484. doi:10.5650/jos.58.481 4. Nishiwaki, N.; Kobiro, K.; Kiyoto, H.; Hirao, S.; Sawayama, J.; Saigo, K.; Okajima, Y.; Uehara, T.; Maki, A.; Ariga, M. Org. Biomol. Chem. 2011, 9, 2832–2839. doi:10.1039/c0ob01005g 5. Nishiwaki, N.; Uehara, T.; Asaka, N.; Tohda, Y.; Ariga, M.; Kanemasa, S. Tetrahedron Lett. 1998, 39, 4851–4852. doi:10.1016/S0040-4039(98)00919-8 6. Nishiwaki, N.; Kobiro, K.; Hirao, S.; Sawayama, J.; Saigo, K.; Ise, Y.; Okajima, Y.; Ariga, M. Org. Biomol. Chem. 2011, 9, 6750–6754. doi:10.1039/c1ob05682d

License and Terms This is an Open Access article under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

7. Nishiwaki, N.; Kobiro, K.; Hirao, S.; Sawayama, J.; Saigo, K.; Ise, Y.; Nishizawa, M.; Ariga, M. Org. Biomol. Chem. 2012, 10, 1987–1991. doi:10.1039/c2ob06925c 8. Balasubramaniam, S.; Aidhen, I. S. Synthesis 2008, 3707–3738. doi:10.1055/s-0028-1083226

The license is subject to the Beilstein Journal of Organic Chemistry terms and conditions: (http://www.beilstein-journals.org/bjoc)

A review and references are cited in. 9. Parhi, A. K.; Franck, R. W. Org. Lett. 2004, 6, 3063–3065. doi:10.1021/ol0489752 Nitrile oxide having a Weinreb amide can be generated from cinnamic acid within five reaction steps; however, this protocol does not involve

The definitive version of this article is the electronic one which can be found at: doi:10.3762/bjoc.11.138

a conversion of N-methylamide to the Weinreb amide. 10. Nagashima, H.; Ozaki, N.; Washiyama, M.; Itoh, K. Tetrahedron Lett. 1985, 26, 657–660. doi:10.1016/S0040-4039(00)89172-8 11. Katritzky, A. R.; Hoffmann, S.; Suzuki, K. ARKIVOC 2004, xii, 14–22. doi:10.3998/ark.5550190.0005.c03 12. Fang, C.; Qian, W.; Bao, W. Synlett 2008, 2529–2531. doi:10.1055/s-2008-1078218 13. Wu, X.; Larhed, M. Org. Lett. 2005, 7, 3327–3329. doi:10.1021/ol0512031 14. Dokli, I.; Gredičak, M. Eur. J. Org. Chem. 2015, 2727–2732. doi:10.1002/ejoc.201500051 15. Zhou, F.; Ding, K.; Cai, Q. Chem. – Eur. J. 2011, 17, 12268–12271. doi:10.1002/chem.201102459

1245

Development of variously functionalized nitrile oxides.

N-Methylated amides (N,4-dimethylbenzamide and N-methylcyclohexanecarboxamide) were systematically subjected to chemical transformations, namely, N-to...
NAN Sizes 0 Downloads 10 Views